Solution Properties of a Fluorinated Alkyl Methacrylate Polymer in

Ji Guo, Pascal André, Mireille Adam, Sergey Panyukov, Michael Rubinstein*, and Joseph M. DeSimone*. Department of Chemistry, University of North Caro...
0 downloads 0 Views 125KB Size
Macromolecules 2006, 39, 3427-3434

3427

Solution Properties of a Fluorinated Alkyl Methacrylate Polymer in Carbon Dioxide Ji Guo,† Pascal Andre´ ,† Mireille Adam,† Sergey Panyukov,†,§ Michael Rubinstein,*,† and Joseph M. DeSimone*,†,‡ Department of Chemistry, UniVersity of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599-3290; Department of Chemical and Biomolecular Engineering, North Carolina State UniVersity, Raleigh, North Carolina 27695-7905; and P. N. LebedeV Physics Institute, Russian Academy of Sciences, Moscow 117924, Russia ReceiVed NoVember 11, 2005; ReVised Manuscript ReceiVed March 8, 2006

ABSTRACT: The solution properties of a fluorinated alkyl methacrylate, poly(1,1,2,2-tetrahydroperfluorooctyl methacrylate) (PFOMA), in liquid and supercritical carbon dioxide (CO2) were studied by static and dynamic light scattering. The solvent quality of CO2 was found to improve with increasing temperature and CO2 density as exhibited by an increase of the second virial coefficient. Both the hydrodynamic radius expansion factor and the second virial coefficient of PFOMA solution were found to be functions of a single interaction parameter that can be independently changed by either temperature or density variations. Furthermore, we demonstrate that the relationship between two directly measurable quantities, the second virial coefficient and the hydrodynamic expansion ratio, is the same for both temperature-induced and CO2 density-induced variations of solvent quality.

I. Introduction As a solvent, supercritical carbon dioxide (scCO2) possesses many unique characteristics including an easily accessible critical point1 (a critical temperature, Tc, of 31.1 °C and a critical pressure, pc, of 73.8 bar).2 A unique feature of scCO2 is the ability to easily tune the solvent quality by changing the temperature (T) or CO2 density (F).3 Many amorphous fluoropolymers are soluble in liquid or supercritical carbon dioxide.1,4 The ability to dissolve polymers in scCO2 creates new opportunities in chemical manufacturing, such as spin-coating and spray-coating from liquid CO2, separations, and complexation of organic acids and heavy metals.5-13 Therefore, it is necessary to understand the solvent properties of CO2 for the various polymers envisaged for these applications. Various experimental techniques have been used to characterize polymer solutions in carbon dioxide.14-19 The most commonly used method to characterize polymer solubility in a supercritical fluid is to perform phase equilibrium measurements to determine cloud point curves as a function of temperature or CO2 density at a given polymer concentration.20 Recently, highpressure scattering methods have been used for quantitative measurements of polymer solution behavior to get information about interactions and molecular sizes in carbon dioxide.21-23 In this work, we study the influence of solvent density and temperature on the properties of a fluorinated poly(alkyl methacrylate), PFOMA, in CO2. We investigate the change of solvent quality using two different approaches, as shown in Figure 1: (1) by varying the temperature at a constant CO2 density; (2) by changing CO2 density at a constant temperature. Using high-pressure static and dynamic light scattering, we measured the molecular weights and sizes of fractionated PFOMA samples. The second virial coefficient and hydrodynamic radius were found to increase with increasing temperature †

University of North Carolina at Chapel Hill. North Carolina State University. § Russian Academy of Sciences. * To whom correspondence should be addressed: e-mail [email protected], [email protected]. ‡

Figure 1. Theta curve (dependence of θ-density on θ-temperature) for PFOMA solution in CO2.

and CO2 density. The focus of this work was to construct functions of a single interaction parameter which enable one to quantitatively predict the strength of interactions and sizes of PFOMA in CO2 with the change of temperature or solvent density for different PFOMA molecular weights. For example, one can use this function to calculate how to tune the temperature and CO2 density to get a desired polymer size or interaction parameter. We demonstrate that both the second virial coefficient and the hydrodynamic radius expansion coefficient can be expressed as functions of a single interaction parameter z that varies with temperature (T) at constant CO2 density (F) as N1/2[1 - θ(F)/T] and with CO2 density at constant temperature as N1/2[1 - Fθ(T)/F], where θ(F) is the theta temperature for given density F, Fθ is the theta density (the CO2 density at the theta condition), and N is the number of Kuhn segments of the polymer chains. The relationship between the second virial coefficient (A2) and the hydrodynamic radius expansion factor (Rh) was found to follow the same behavior (independent of directions in Figure 1). Therefore, we verified that two different ways of varying the interaction parameter z (directions (1) and (2) in Figure 1) are directly related and found that the θ-temperature varies reciprocally proportional to CO2 density

10.1021/ma052409k CCC: $33.50 © 2006 American Chemical Society Published on Web 04/08/2006

3428

Guo et al.

Macromolecules, Vol. 39, No. 9, 2006

θ(F) ) (260/F)(KmL/g), while the θ-density varies reciprocally proportional to the absolute temperature Fθ(T) ) (260/T)(g K/mL). Thus, we have confirmed the existence of a single interaction parameter that combines the temperature, the solvent density, and the degree of polymerization into a single variable. II. Experimental Section II.1. Materials. The monomer 1,1,2,2-tetrahydroperfluorooctyl methacrylate (FOMA) (provided by DuPont) was purified and deinhibited by passing it through an alumina column. The initiator 2,2′-azobis(isobutyronitrile) (AIBN, Kodak, 99%) was recrystallized twice in methanol (Aldrich). All purification solvents were purchased from Aldrich and used as received. II.2. Synthesis of PFOMA. FOMA monomers were purged with argon for ∼15 min prior to transferring into a 25 mL high-pressure view cell containing AIBN (0.5-1 wt %) and a magnetic stirring bar. The contents of the high-pressure cell were purged with argon for additional 15 min, and then the reaction cell was heated to 60 °C while CO2 was added via syringe pump (Isco) over ca. 15 min of time to a pressure of 345 bar. The polymerization was continued for 24 h at 60 °C and 345 bar. The resulting polymer and any unreacted monomer were removed from the reaction cell by dissolving all of the contents in 1,1,2-trifluorotrichloroethane. The polymer was precipitated into a large excess of methanol, isolated by suction filtration, washed several times with methanol, and dried in a vacuum oven overnight under reduced pressure. II.3. Fractionation and Preparation of Polymer Solutions. The isolated PFOMA was fractionated to reduce the polydispersity of samples used in the analysis. Under isothermal conditions, PFOMA was fractionated by applying an increasing CO2 pressure profile.3 The fractionation temperature was 60 °C, and the pressure was increased from 106 to 414 bar with a step interval of 14 bar. A total of 13 fractions were isolated at different CO2 densities. Each fraction was 1.5-2 g in mass and was numbered consecutively from 1 to 13. Fractions 2 and 8 were used to measure the refractive index, and fractions 3 and 6 were used to do all other measurements. The concentration of the polymer solutions used in the light scattering experiments ranged from 0.01 to 0.04 g/mL. The polymer samples were weighed and sealed into a high-pressure optical cell (see details in the Supporting Information). The cell volume was adjusted using a piston at the top of the optical cell. Carbon dioxide was filtered and injected slowly into the cell at room temperature, until the initial pressure reached 131 bar. The polymer solution was then heated to 60 °C with constant stirring until it became completely transparent. Subsequently, the solution was stabilized without stirring for 1.5 h while the transmission intensity of the polymer solution was detected using a LaserPAD power meter (Coherent Inc.). All experimental conditions were reached by decreasing temperatures and/or by increasing the CO2 density. The scattering intensity was monitored by a photometer with each temperature and density change until the intensity became stable. II.4. Measurement of Refractive Index Increment. The refractive index increment (dn/dc) quantifies how the refractive index of the polymer solution changes with polymer concentration. For traditional solvents, dn/dc is measured with a differential refractometer, and the values for many systems are available in the literature.24 The values of dn/dc for most polymers in CO2, however, are unknown and needed to be measured directly. The measurements of the refractive index increment of two fractions of PFOMA in CO2 are described in the Supporting Information. II.5. Light Scattering. Static light scattering (SLS) and dynamic light scattering (DLS) experiments were performed by means of a spectrometer equipped with an argon laser (see the Supporting Information). The scattering wave vector q is given by q)

4πns sin(Θ/2) λ

(1)

where ns is the refractive index of the solvent, λ is the wavelength of the light in the vacuum, and Θ is the scattering angle.25 The

Table 1. Rayleigh Ratios RθCO2 of CO2 at Scattering Angle 90° for Different CO2 Densities at 25 °C CO2 density (g/mL) RθCO2 × 105 (cm-1)

0.86 3.87

0.89 3.37

0.93 2.75

0.97 2.26

1.01 1.97

refractive index values ns for carbon dioxide were measured under experimental conditions. The range of the external detection angle was varied from 25° to 155°, which results in a scattering wavevector q that ranged from 0.010 to 0.027 nm-1. All DLS experiments were performed at angles 40°, 90°, and 140°. II.5.1. Static Light Scattering. Static light scattering experiments measure the relative excess of scattering intensity with respect to the solvent I(q,c) for various polymer concentrations and I(q,c) ) (Is - ICO2)/ICO2, where the Is and ICO2 are the scattering from the polymer solution and from CO2, respectively. The scattering data were converted into the excess Rayleigh’s ratio, R(q,c), using the 2 equation26 R(q,c) ) I(q,c)RCO ) I(q,c)Rtoluene ICO2/Itoluene. The θ θ toluene Rayleigh ratio of toluene is Rθ ) 3.21 × 10-5 cm-1 at 25 °C 2 for λ ) 514 nm, and the Rayleigh ratios of CO2 (RCO θ ) at 25 °C are listed in Table 1. The dependence of the excess Rayleigh ratio R(q,c) on scattering wavevector q and polymer concentration c is expressed as26 1 Kc 1 1 + Rg2q2... ) 3 R(q,c)0) Mw

(

)

(2)

Kc 1 (1 + 2MwA2c...) ) R(q)0,c) Mw

(3)

Here A2 is the second virial coefficient, Rg is the z-average radius of gyration, and Mw is the weight-average molecular weight. The optical constant K is defined by26 K)

4π2ns2 dn λ4N dc av

2

( )

(4)

where Nav is Avogadro’s number. II.5.2. Dynamic Light Scattering. Dynamic light scattering (DLS) measures the autocorrelation function g(2)(q,t) of the scattering intensity.26 The autocorrelation function depends on how molecules move on the length scale 1/q, with a characteristic time τ)

1 Dq2

(5)

where D is the diffusion coefficient. For dilute solutions, the concentration dependence of the diffusion coefficient D(c) can be approximated as27 D(c) ) D0(1 + kDc)

(6)

where kD is the diffusional second virial coefficient and D0 is the diffusion coefficient at infinite dilution D0 )

kBT 6πηsRh

(7)

where ηs is the solvent viscosity, kB is the Boltzmann constant, T is the absolute temperature, and Rh is the hydrodynamic radius. Equation 7 is used to calculate hydrodynamic radius Rh from the diffusion coefficient D0 obtained by extrapolation of D(c) (eq 6) to infinite dilution (c f 0). II.6. Interaction Parameter. The second virial coefficient of polymers dissolved in organic solvents can be collapsed onto a single curve without any adjustable parameters as was first demonstrated by Berry for polystyrene solutions.28 In a good solvent

Fluorinated Alkyl Methacrylate Polymer in CO2 3429

Macromolecules, Vol. 39, No. 9, 2006 Table 2. Mass of Kuhn Segment Ms and Kuhn Length b of PFOMA, PMMA,30 and PS28 polymer

Ms (Da)

b (nm)

PFOMA PMMA PS

6300 666 728

3.0 1.7 1.8

and in a θ-region,29 A2Mw1/2 is only a function of chain interaction parameter, z. A2Mw1/2

M03/2 3

Navb

{

) f(z) ≈ CA2

z 0.528

z

for |z| < 1 (θ region) for z > 1 (good solvent) (8)

where CA2 is a numerical coefficient. Kuhn segment mass, Ms, and length, b, are defined through the equivalent freely jointed chain model.29 b)

Ms ) 6

6Rgθ2 ms l Mw

(9)

( )

(10)

Rgθms 2 1 l Mw

where ms is the molar mass of the monomer, Rgθ is the radius of gyration at the θ-condition, and the contour length of the monomer is l ) 2 × 1.54 Å × sin(68°) ) 2.6 Å. The length b and the mass Ms of the Kuhn segment for PFOMA (see section IV.2), polystyrene (PS),28 and poly(methyl methacrylate) (PMMA)30 are listed in Table 2. The interaction parameter z at constant solvent density can be expressed in terms of the reduced temperature (see direction (1) in Figure 1) z ) CTN1/2

T - θ(F) (F ) const) T

(11)

where CT is a numerical coefficient related to the temperature dependence of interaction parameter z and N is the number of Kuhn segments, N ) Mw/M0. Similarly, the interaction parameter z at constant temperature can be expressed in terms of the reduced density (see direction (2) in Figure 1) z ) CFN1/2

F - Fθ(T) (T ) const) F

(12)

where CF is a numerical coefficient related to the density dependence of the interaction parameter z. In section IV.3 we present a theoretical analysis of the temperature and density dependencies of the second virial coefficient in supercritical solvents and derive the temperature dependence of the theta density Fθ(T). Numerical coefficients CA2, CT, and CF can depend on chemical structure of polymer. For a particular polymer-solvent pair, these coefficients should be independent of molecular weight, and we will demonstrate below that CT and CF are equal to each other for PFOMA in CO2. In the θ-region, the second virial coefficient is linearly proportional to the interaction parameter z (see eq 8 with eq 11 for temperature variations and eq 12 for density variations). 3/2

A2Mw1/2

{

M0

Navb

[

) CA2z ) 3

CFA2N1/2

]

θ(F) T Fθ(T) 1F

CTA2N1/2 1 -

[

]

for temperature variations (CTA2 ) CA2CT) for density variations

(CFA2 ) CA2CF) (13)

Figure 2. (A) Refractive index of carbon dioxide as a function of CO2 density at 25 °C: values calculated from ref 32 (O) and our measurements (b). (B) dn/dc as a function of CO2 density at 25 °C for two PFOMA fractions: fraction 2 (4) and fraction 8 (2).

The hydrodynamic radius expansion factor Rh is defined as the ratio of hydrodynamic radius Rh at a given condition to the hydrodynamic radius at the θ-condition, Rh(θ). This expansion ratio Rh, describing the swelling of the chain, is also a function of a single interaction parameter z27,31 (related to the excluded volume interactions). The hydrodynamic radius expansion factor Rh can be approximated in the θ-region by a linear function of the interaction parameter z.32

{

Rh ) 1 + CRz )

[

1 + CFRN1/2

]

θ(F) T Fθ(T) 1F

1 + CTRN1/2 1 -

[

]

for temperature variations (CTR ) CRCT) for density variations

(CFR ) CRCF) (14)

where CR is a numerical coefficient. Furthermore, the existence of a single interaction parameter for both temperature and density variations implies that these coefficients are not independent, but are related to each other by CTR CTA2

)

CFR CR ) CA2 CFA2

(15)

as can be deduced from eqs 13 and 14.

III. Results III.1. Molecular Weight Mw. The measurements of the refractive index and the static light scattering experiments were carried out in order to determine the molecular weights of the various PFOMA fractions. The values of refractive index of carbon dioxide calculated from the literature25 have been confirmed in our experiments, as shown in Figure 2A. The refractive index values of carbon dioxide were found to increase linearly with CO2 density. The values of the refractive index

3430

Guo et al.

Figure 3. Variation of the second virial coefficient A2 with (A) temperature at constant CO2 density F ) 0.86 g/mL, PFOMA with Mw ) 300 kDa (]) and Mw ) 900 kDa ([); (B) CO2 density at constant temperature T ) 25 °C, PFOMA with Mw ) 300 kDa (0) and Mw ) 900 kDa (9).

increments (dn/dc) at different CO2 densities at 25 °C for two PFOMA fractions (fractions 2 and 8) are displayed in Figure 2B. As the carbon dioxide density increases, the dn/dc values for both fractions of PFOMA in CO2 decrease from 0.12 to 0.10 mL/g. The effect of CO2 density on the dn/dc of the polymer solutions can be mostly attributed to the change of CO2 refractive index with CO2 density (see Supporting Information). The dn/dc values are independent of PFOMA molecular weight, as expected for long polymeric chains where the end groups have negligible effects on the polymer solution properties. The data for fractions 2 and 8 at the same CO2 density agree with each other within experimental accuracy. The molecular weights of fractions 3 and 6 were measured by light scattering using dn/dc values determined from fractions 2 and 8. The molecular weights of fractions 3 and 6 were found to be 300 ( 30 and 900 ( 70 kDa, respectively. III.2. Second Virial Coefficient A2. The solvent quality change can be monitored through measurements of A2 at different temperatures and solvent densities of CO2. The second virial coefficient, A2, was determined from the concentration dependence of the scattering intensity from dilute PFOMA solutions. The temperatures and densities corresponding to θ-conditions with the second virial coefficient equal to zero can be determined at constant density by varying the temperature (direction (1) in Figure 1) or at constant temperature by varying the density (direction (2) in Figure 1). Previous small-angle neutron scattering experiments have demonstrated that not only does a θ-temperature but a θ-density can also exist for a CO2polymer solution.33 When the CO2 density was kept constant (0.86 g/mL), A2 was found to increase with increasing temperature (Figure 3A),

Macromolecules, Vol. 39, No. 9, 2006

changing from negative to positive values. In the range of experimental conditions investigated, the PFOMA samples with both molecular weights (300 and 900 kDa) had the same θ-point [θ ) 27 ( 1 °C, Fθ ) 0.86 g/mL]. The variation of A2 with carbon dioxide density at constant temperature (25 °C) is displayed in Figure 3B. A2 was found to increase with increasing density, ranging from negative to positive values. The increase of the second virial coefficient with density is stronger for the lower molecular weight fraction. The θ-condition was determined to be [θ ) 25 °C, Fθ ) 0.88 ( 0.02 g/mL]. The improvement of the solvent quality quantitatively assessed by static light scattering was confirmed by the variation of the diffusional second virial coefficient kd deduced from the dynamic light scattering data (see Supporting Information). III.3. Size Variations of Polymer Chains. The radius of gyration of the fraction with Mw ) 900 kDa is Rg ) 15 ( 4 nm, at θ-conditions [θ ) 27 (1 °C, F(θ) ) 0.86 g/mL] and [θ ) 25 °C, F(θ) ) 0.88 ( 0.02 g/mL]. There was no measurable change of Rg within the experimental error with variations of either CO2 density or temperature (see Figure SI-4 in Supporting Information). For the fraction with Mw ) 300 kDa, the radius of gyration could not be measured by static light scattering. The values of Rg for PFOMA with Mw ) 900 kDa are close to the detection limit of our instrument. The difficulty of the measurements of Rg for the high molecular weight sample is clearly displayed in the plot of c/I(q,c) ∼ q2 (see Figure SI-3B in the Supporting Information). The contrast between the polymer and supercritical CO2 is much lower than the contrast in organic systems. The hydrodynamic radius Rh was measured by the dynamic light scattering. The hydrodynamic radii Rh at θ-conditions [θ ) 27 (1 °C, Fθ ) 0.86 g/mL] and [θ ) 25 °C, Fθ ) 0.88 ( 0.02 g/mL] are 6.4 ( 0.1 and 11.4 ( 0.2 nm for the low (300 kDa) and the high (900 kDa) molecular weight fractions (fractions 3 and 6), respectively. The ratio of Rg/Rh for the sample with molecular weight 900 kDa was found to be 1.3 ( 0.4 at θ-conditions, which is very close to the value reported for monodisperse linear polymers.29 Figure 4 illustrates the dependence of hydrodynamic radius expansion ratio Rh on temperature (A) and CO2 density (B). The values of Rh increase with increasing temperature and CO2 density. This observation verifies that the polymer size increases with the improvement of solvent quality along both the temperature and density directions (see Figure 1). The increase of hydrodynamic radius expansion factor with temperature or density is stronger for the higher molecular weight fraction. IV. Discussion IV.1. Hydrodynamic Radius. It is demonstrated in Figure 5A that the hydrodynamic radius expansion factor for different polymer solutions can be expressed as the universal function of the reduced temperature N1/2[1 - θ(F)/T] (eq 14), including the data for PFOMA in CO2 and the data from earlier experiments on polystyrene (PS) in two different solvents.32,34 In the θ-region, Rh is approximately a linear function of reduced temperature N1/2[1 - θ(F)/T], and the numerical coefficient CTR is 0.06 ( 0.01 for both PFOMA in CO2 and polystyrene in organic solvents.32,34,35 The radius of gyration expansion factor Rg (defined as the ratio of Rg at a given condition to the radius of gyration Rgθ at the θ-point) is also a universal function of reduced temperature N1/2[1 - θ(F)/T].28,36 The earlier experimental results for PS32,34 and PMMA37 in different organic solvents also exhibit the linear relationship between Rg and reduced temperature in the θ-region that can be written as Rg

Macromolecules, Vol. 39, No. 9, 2006

Figure 4. Hydrodynamic expansion factor Rh: (A) as a function of temperature at constant CO2 density of F ) 0.86 g/mL, PFOMA with Mw ) 300 kDa (]) and Mw ) 900 kDa ([); (B) as a function of CO2 density at constant temperature T ) 25 °C. PFOMA with Mw ) 300 kDa (0) and Mw ) 900 kDa (9).

) 1 + (0.08 ( 0.02)N1/2(1 - θ/T) (the inset in Figure 5A). The coefficients CTR for hydrodynamic radius expansion factor Rh and for the radius of gyration expansion factor Rg of PS in trans-decalin are equal to each other within experimental error, as expected from the fact that the ratio of Rg/Rh should not have significant temperature dependence in the θ-region. It is interesting to note that numerical coefficients CTR for different polymers (PFOMA, PS, and PMMA) are close to each other. We observe for the first time that polymer size is a function of the reduced solvent density N1/2[1 - Fθ(T)/F] (direction (2) in Figure 1). There is a linear relationship between the hydrodynamic radius expansion factor Rh and the reduced density in the θ-region (see Figure 5B and eq 14), and the numerical coefficient CFR ) 0.06 ( 0.01 turned out to be equal to CTR within the experimental accuracy. IV.2. Second Virial Coefficient. It was reported that experimental data as well as computer simulation data for the second virial coefficient of different polymers (e.g., polystyrene28 and poly(methyl methacrylate)30) in various organic solvents can be expressed as functions of the interaction parameter z ∼ N1/2(1 - θ/T) (see eq 13). The dependence of the normalized second virial coefficient on the reduced temperature N1/2[1 - θ(F)/T] of PFOMA solutions in CO2 is presented in Figure 6A. The data of PFOMA samples for both molecular weights collapse onto a single curve. Most of the PFOMA points are in θ-region, where the interaction parameter is small and the dependence of normalized second virial coefficient on reduced temperature N1/2[1 - θ(F)/T] is linear. The numerical coefficient for the θ-region of PFOMA solution in CO2 is CTA2 ) 0.50 ( 0.05 (eq 13), which is larger than the coefficients 0.39 ( 0.04 for PS and PMMA in organic solvents (shown in the inset in Figure 6A). This difference in coefficients is probably due to different chemical structures of the poly-

Fluorinated Alkyl Methacrylate Polymer in CO2 3431

Figure 5. (A) Hydrodynamic radius expansion factor Rh as a function of the reduced temperature N1/2[1 - θ(F)/T] at constant CO2 density F ) 0.86 g/mL with θ(F) ) 27 °C, for PFOMA with Mw ) 300 kDa (]) and Mw ) 900 kDa ([), PS in cyclohexane32 (3), PS in trans-decalin34 (1). Inset: radius of gyration expansion factor Rg ) Rg/Rg(θ) as a function of the chain interaction parameter N1/2(1 - θ/T), for PS in cyclohexane32 (3), PS in trans-decalin34 (1), PMMA in water and tertbutyl alcohol37 (4). (B) Hydrodynamic radius expansion ratio Rh as a function of the reduced density N1/2[1 - Fθ(T)/F] at constant temperature T ) 25 °C with Fθ ) 0.88 g/mL, for PFOMA with Mw ) 300 kDa (0) and Mw ) 900 kDa (9).

mers: the PFOMA has larger Kuhn segment than PS and PMMA (see Table 1). In a compressible solvent like CO2, the interaction parameter can be easily varied by changing the solvent density. The behavior of the second virial coefficients as a function of reduced density N1/2[1 - Fθ(T)/F] is presented in Figure 6B. The data for two different molecular weight fractions of PFOMA collapse onto a single curve similar to the curve in Figure 6A. This is the first demonstration that the second virial coefficient A2 of polymers in compressible solvents is a function of the single interaction parameter z ∼ N1/2[1 - Fθ(T)/F]. The numerical coefficient of the linear dependence of the normalized second virial coefficient on the reduced density N1/2[1 Fθ(T)/F] for the θ-region is CFA2 ) 0.50 ( 0.05 (eq 13), which is equal to the value of coefficient CTA2 within experimental error. The ratios of numerical coefficients for the temperature variation (CTR/CTA2 ) CR/CA2 = 0.12) and for the density variation (CFR/CFA2 ) CR/CA2 = 0.12) are equal to each other, as expected from eq 15. The plots in Figures 5 and 6 are based on assumptions that the interaction parameter is a linear function of either reduced temperature or reduced density. To combine solvent quality changes with both temperature and density variations (both directions in Figure 1), the dependence of hydrodynamic radius

3432

Guo et al.

Macromolecules, Vol. 39, No. 9, 2006

sion factor of PFOMA in CO2 can be approximated in the θ-region by a linear function of the second virial coefficient.

Rh ) 1 +

M 3/2 CR 1/2 0 A M ) 2 w C A2 Navb3 1 + (0.12 ( 0.02)A2Mw

1/2

M03/2 Navb3

(16)

The data of PS and PMMA also collapse onto a single curve but with a higher slope of 0.2 ( 0.03. These curves provide a more direct way of predicting the interactions and polymer size in solutions, either getting size information from interactions or obtaining interactions information from polymer size, since both A2 and Rh can be measured directly and independently by the light scattering experiments. IV.3. Theta Curve. We independently verified that numerical coefficients CTR and CFR are equal to each other (see Figure 5) and numerical coefficients CTA2 and CFA2 are equal to each other (see Figure 6). Thus, we have two independent verifications that coefficients CT and CF in eqs 11 and 12 are the same, since these two equations are two different representations of the same interaction parameter z. We conclude that θFθ ) constant and the density dependence of θ-temperature is

θ(F) )

260 gK F mL

( )

(17)

while the temperature dependence of θ-density is Figure 6. (A) Plot of A2M1/2Ms3/2NAV-1b-3 as a function of reduced temperature N1/2[1 - θ(F)/T] at constant CO2 density FCO2 ) 0.86 g/mL with θ(F) ) 27 °C, of PFOMA with Mw ) 300 kDa (]) and Mw ) 900 kDa ([). Inset: the plot of A2M1/2Ms3/2NAV-1b-3 as a function of reduced temperature N1/2(1 - θ/T) for PMMA in 46.8% butanol/53.2% 2-propanol30 (3) and PS in decalin28 (1). (B) Plot of A2M1/2Ms3/2NAV-1b-3 as a function of reduced density N1/2[1 - Fθ(T)/F] at constant temperature T ) 25 °C with Fθ ) 0.88 g/mL, PFOMA with Mw ) 300 kDa (0) and Mw ) 900 kDa (9).

Fθ(T) )

260 KmL T g

( )

(18)

Below we present a theoretical justification of the hyperbolic CO2 density dependence of θ-temperature (Figure 1). In highly compressible solvents (such as CO2 near its critical point), the second virial coefficient of interaction between solvent molecules is small, and the free energy density of the solvent is dominated by the third virial term

FCO2 ≈

B 3F 3 3

(19)

where B3 is solvent interaction parameter. There are additional interaction terms in free energy density of a polymer solution with Kuhn segment number density ck in this compressible solvent: a term proportional to FCO2ck due to polymer-solvent interaction and a term proportional to ck2 due to direct polymerpolymer interaction

F≈

Figure 7. Plot of the hydrodynamic radius expansion factor Rh as a function of A2M1/2Ms3/2NAV-1b-3: PFOMA at constant temperature T ) 25 °C with Fθ ) 0.88 g/mL, of PFOMA with Mw ) 300 kDa (0) and Mw ) 900 kDa (9); PFOMA at constant CO2 density FCO2 ) 0.86 g/mL with θ(F) ) 27 °C, of PFOMA with Mw ) 300 kDa (]) and Mw ) 900 kDa ([).

expansion factor Rh on the second virial coefficient A2M1/2Ms3/2Nav-1b-3 has been plotted in Figure 7 for PFOMA at different CO2 densities and temperatures. In the experimental range, the PFOMA data collapsed onto a single curve with the slope CR/CA2 ) 0.12 ( 0.02. The hydrodynamic radius expan-

ck2 B3F3 + B2Fck + kBTν 3 2

(20)

where B2 is the polymer-solvent interaction parameter, kB is the Boltzmann constant, and ν is the excluded volume of monomers. Because of small interval of temperature variations, we neglect temperature dependences of the interaction parameters B2 and B3. To take into account the density fluctuation of a compressible solvent, we can write the local density F as a sum of an average density Fj and a density fluctuation δF

F ) Fj + δF

(21)

Substituting this expression into free energy density (eq 20) and expanding it to quadratic terms in density fluctuation, we find

Macromolecules, Vol. 39, No. 9, 2006

F≈

Fluorinated Alkyl Methacrylate Polymer in CO2 3433

B3Fj3 + B3Fj2δF + B3Fj(δF)2 + B2Fjck + B2ckδF + 3 ck2 (22) kBTν 2

The sum of density fluctuations over the solution volume is zero, ∫VδF dV ) 0, and therefore the second term in eq 22 vanishes. Note that linear term in solvent density fluctuation B2ckδF does not vanish because the polymer concentration ck fluctuates in a correlated way with solvent density. Minimizing the free energy density

F≈

ck2 B3Fj3 + B3Fj(δF)2 + B2Fjck + B2ckδF + kBTν (23) 3 2

with respect to the solvent density fluctuation δF

∂F ) 2B3FjδF + B2ck ) 0 ∂(δF)

(24)

we find the optimal solvent density fluctuation

δF ) -

B2ck 2B3Fj

(25)

Substituting it back into eq 22, we obtain the free energy density

F≈

(

)

ck2 B22 B3Fj3 kBTν ≈ + B2Fjck + 3 2 2B3Fj B3Fj A2 + B2Fjck + RT c2 (26) 3 2 3

where R ) kBNav is the gas constant. The solution concentration c can be expressed in terms of the number density of Kuhn segment ck as

c)

M0 c Nav k

A2 ≈

M02

(

) ( )

B22 νNav θ 1) 12 2kBTνB3Fj T M0

(28)

const Fj

M0 CR Rh ) 1 + A A2Mw1/2 2 C Navb3 Note that coefficients CR/CA2 ) 0.12 ( 0.02 for PFOMA in CO2 and 0.2 ( 0.03 for PS and PMMA solutions. The nonuniversality of the ratio of coefficients CR/CA2 indicates the existence of additional physical interactions that were not considered in existing polymer models. These interactions depend on chemical structure of polymers and will be discussed in our future publications.38

(29)

Thus, we derived that θ-temperature is inversely proportional to the average solvent density

θ)

We have investigated the solution properties of a fluorinated poly(alkyl methacrylate) using both static and dynamic light scattering. The solvent quality of CO2 was shown to improve with increasing temperature or CO2 density. This trend was confirmed by the evaluation of the second virial coefficient A2 measured by static light scattering. The θ-conditions were found as [θ ) 27 (1 °C, Fθ ) 0.86 g/mL] and [θ ) 25 °C, Fθ ) 0.88 ( 0.02 g/mL]. Both the hydrodynamic radius expansion factor and the normalized second virial coefficient of PFOMA dissolved in CO2 are verified to be functions of the interaction parameter z, which is proportional to N1/2[1 - θ(F)/T] (see eq 11), which is in excellent agreement with the data for other polymers dissolved in traditional organic solvents. We have for the first time demonstrated that the hydrodynamic radius expansion factor Rh and the normalized second virial coefficient are also functions of the same interaction parameter z and are proportional to the reduced solvent density N1/2[1 - Fθ(T)/F] at constant temperature. We have determined and theoretically justified reciprocal dependence of the theta temperature on CO2 density for PFOMA as θ(F) ) (260/F)(g K/mL). These results provide an important explanation of the observation that the solvent quality can be tuned by not only the solvent temperature but also the solvent density in a universal way. We have also established that the hydrodynamic radius expansion factor can be expressed as a single function of the normalized second virial coefficient, independently of the methods used to vary the solvent quality. 3/2

that vanishes at the theta temperature

B22 θ) 2kBνB3Fj

V. Conclusions

(27)

Thus, polymer-solvent interactions lead to an effective second virial coefficient in the θ-region

νNav

in question. The effective diameter d and the Kuhn length b for PS are 8 Å and 18 Å, respectively. The effective diameter of the PFOMA is calculated to be 10 Å using the mass of Kuhn segment Ms, Kuhn length b (Table 2), and polymer density 2 g/cm3. The ratio of d/b for PS and PFOMA is 0.44 and 0.33, and we find that (d/b)PFOMA/(d/b)PS ) 0.75. This simple theoretical estimate is in good agreement with experimentally observed ratio(CTA2)PFOMA/(CTA2)PS ) 0.78 ( 0.07.

Acknowledgment. We thank the Kenan Center for the Utilization of Carbon Dioxide in Manufacturing and the NSF Science and Technology Center for Environmentally Responsible Solvents and Process (No. CHE-9876674) for financial support.

(30)

We have verified this dependence and found the value of the constant 260 g K/mL (eq 17) in our experiments. Substituting the second virial coefficient A2 (eq 28) into eq 13, and using an estimation ν ≈ db2 for the excluded volume of the rod of the length b and the effective diameter d, we find that CTA2 ≈ ν/b3 ≈ d/b. Therefore, we conclude that the coefficient CTA2 is not universal, and it depends on chemical structure of the polymer

Supporting Information Available: Experimental details. This material is available free of charge via the Internet at http:// pubs.acs.org.

References and Notes (1) DeSimone, J. M.; Guan, Z.; Elsbernd, C. S. Science 1992, 257, 945947. (2) Quinn, E. L.; Jones, C. L. Carbon Dioxide; Reinhold Publishing Corp.: New York, 1936.

3434

Guo et al.

(3) McHugh, M. A.; Krukonis, V. J. Supercritical Fluid Extraction: Principles and Practice, 2nd ed.; Butterworth-Heineman: Stoneham, 1993. (4) O’Neill, M. L.; Cao, Q.; Fang, R.; Johnston, K. P.; Wilkinson, S. P.; Smith, C. D.; Kerschner, J. L.; Jureller, S. H. Ind. Eng. Chem. Res. 1998, 37, 3067-3079. (5) Garg, A.; Gulari, E.; Manke, C. W. Marcromolecules 1994, 27, 56435653. (6) Tomasko, D. L.; Li, H.; Liu, D.; Han, X.; Wingert, M. J.; Lee, L. J.; Koelling, K. W. Ind. Eng. Chem. Res. 2003, 42, 6431-6456. (7) DeSimone, J. M.; Maury, E. E.; Menceloglu, Y. Z.; Mcclain, J. B.; Romack, T. J.; Combes, J. R. Science 1994, 265, 356-359. (8) Adamsky, F. A.; Beckman, E. J. Macromolecules 1994, 27, 312314. (9) Dillow, A. K.; Yun, S. L.; Suleiman, D. S.; Boatright, D. L.; Liotta, C. L.; Eckert, C. A. Ind. Eng. Chem. Res. 1996, 35, 1801-1806. (10) Johnston, K. P.; Harrison, K. L.; Clarke, M. J.; Howdle, S. M.; Heitz, M. P.; Bright, F. V.; Carlier, C.; Randolph, T. W. Science 1996, 271, 624-626. (11) Randolph, T. W. Trends Biotechnol. 1990, 8, 78. (12) McAllister, K.; Sazani, P.; Adam, M.; Cho, M. J.; Rubinstein, M.; Samulski, R. J.; DeSimone, J. M. J. Am. Chem. Soc. 2002, 124, 15198-15207. (13) Lemert, R. M.; Johnston, K. P. Ind. Eng. Chem. Res. 1991, 30, 1222. (14) Mertsch, R.; Wolf, B. A. Macromolecules 1994, 27, 3289-3294. (15) Cotts, P. M. Macromolecules 1994, 27, 6487-6491. (16) Fulton, J. L.; Pfund, D. M.; McClain, J. B.; Romack, T. J.; Maury, E. E.; Combes, J. R.; Samulski, E. T.; DeSimone, J. M.; Capel, M. Langmuir 1995, 11, 4241-4249. (17) Guan, Z. Department of Chemistry, University of North Carolina, Chapel Hill, 1994. (18) Chu, B.; Lin, J. S. J. Chem. Phys. 1970, 53, 4454. (19) Nagashima, K.; Lee, C. T.; Xu, B.; Johnston, K. P.; DeSimone, J. M.; Johnson, C. S. J. Phys. Chem. B 2003, 107, 1962-1968. (20) Folk, S. L. Department of Chemistry, University of North Carolina, Chapel Hill, 2003. (21) Buhler, E.; Dobrynin, A. V.; DeSimone, J. M.; Rubinstein, M. Macromolecules 1998, 31, 7347-7355. McClain, J. B.; Londono, D.;

Macromolecules, Vol. 39, No. 9, 2006

(22) (23) (24) (25) (26) (27) (28) (29) (30) (31) (32) (33) (34) (35) (36) (37) (38) (39)

Combes, J. R.; Romack, T. J.; Canelas, D. A.; Betts, D. E.; Wignall, G. D.; Samulski, E. T.; DeSimone, J. M. J. Am. Chem. Soc. 1996, 118, 917-918. Wignall, G. D.; Melnichenko, Y. B.; Triolo, R.; McClain, J. B.; DeSimone, J. M. Abstr. Pap. Am. Chem. Soc. 2001, 221, U374-U375. ChilluraMartino, D.; Triolo, R.; McClain, J. B.; Combes, J. R.; Betts, D. E.; Canelas, D. A.; DeSimone, J. M.; Samulski, E. T.; Cochran, H. D.; Londono, J. D. J. Mol. Struct. 1996, 383, 3-10. Huglin, M. B. J. Appl. Polym. Sci. 1965, 9, 3963-4001, 4003-4024. Burns, R. C.; Graham, C.; Weller, A. R. M. Mol. Phys. 1986, 59, 41-64. Chu, B. Laser Light Scattering: Basic Principles and Practice, 2nd ed.; Academic Press: Boston, 1991. Yamakawa, H. Modern Theory of Polymer Solutions; Harper and Row: New York, 1970. Berry, G. C. J. Chem. Phys. 1966, 44, 4550-4564. Rubinstein, M.; Colby, R. H. Polymer Physics; Oxford University Press: New York, 2003. Schulz, G. V.; Inagaki, H.; Kirste, R. Z. Phys. Chem. (Munich) 1960, 24, 290-404. Arai, T.; Abe, F.; Yoshizaki, T.; Einaga, Y.; Yamakawa, H. Macromolecules 1995, 28, 3609-3616. Arai, T.; Abe, F.; Yoshizaki, T.; Einaga, Y.; Yamakawa, H. Macromolecules 1995, 28, 5458-5464. Melnichenko, Y. B.; Kiran, E.; Heath, K. D.; Salaniwal, S.; Cochran, H. D.; Stamm, M.; Van Hook, W. A.; Wignall, G. D. J. Appl. Crystallogr. 2000, 33, 682-685. Nose, T.; Chu, B. Macromolecules 1979, 12, 1122-1129. Chu, B.; Nose, T. Macromolecules 1980, 13, 122-132. Konishi, T.; Yoshizaki, T.; Saito, T.; Einaga, Y.; Yamakawa, H. Macromolecules 1990, 23, 290-297. Mitsuo, N. Phys. ReV. E 1995, 51, 5770. Panyukov, S.; Shirvanyants, D.; Liao, Q.; Rubinstein, M., to be published. Akcasu, A. Z. Polymer 1981, 22, 1169-1180.

MA052409K